Ion Mobility Unlocks the Photofragmentation ... - ACS Publications

Sep 2, 2014 - ... are analyzed in an ion mobility mass spectrometer (IMMS) and found to .... S. Glover , U. Rothlisberger , David E. Clemmer , Thomas ...
60 downloads 0 Views 1MB Size
Letter pubs.acs.org/JPCL

Ion Mobility Unlocks the Photofragmentation Mechanism of Retinal Protonated Schiff Base N. J. A. Coughlan, B. D. Adamson, K. J. Catani, U. Wille, and E. J. Bieske* School of Chemistry, The University of Melbourne, Melbourne, Victoria, Australia 3010 S Supporting Information *

ABSTRACT: Retinal protonated Schiff base (RPSB) is a key molecular component of biological photoreceptors and bacterial photosynthetic structures, where its action involves photoisomerization around bonds in the polyene chain. In a vacuum environment, collisional activation or exposure to visible light causes the RPSB molecule to disintegrate, producing charged molecular fragments with m/z = 248 Da that cannot be formed by simple cleavage of the polyene chain. Photofragments resulting from laser excitation of RPSB at a wavelength of 532 nm are analyzed in an ion mobility mass spectrometer (IMMS) and found to be the protonated Schiff base of β-ionone. Density functional theory calculations at the M06-2X/cc-pVDZ level support a fragmentation mechanism in which RPSB undergoes an electrocyclization/ fragmentation cascade with the production of protonated Schiff base of β-ionone and toluene. SECTION: Spectroscopy, Photochemistry, and Excited States

R

photofragments, which have different structures (and collision cross sections) for the two mechanisms. IMMS is a versatile tool for exploring conformations of charged molecules, relying on the sensitivity of a molecule’s collision cross section to its structure; under the influence of an electric field, compact molecules travel more swiftly through a buffer gas than extended, unfolded molecules.8−13 IMMS has been employed to elucidate the structures of a broad range of charged molecules and clusters including carbon and silicon clusters,14,15 carbohydrates,16 peptides,17,18 and proteins.10 Recently, we showed that it can also be used to monitor the photoisomerization of charged molecules in the gas phase including carbocyanine dyes and RPSB.19−21 Here, we demonstrate that IMMS can be used to explore the structures of molecular photofragments and shed light on photofragmentation mechanisms. In our experiment, RPSB ions produced by electrospray ionization are irradiated in the drift region of a purpose-built ion mobility spectrometer in which ions are propelled through N2 buffer gas by a modest electric field (Figure 2).19−21 This arrangement enables us to discriminate charged photofragments on the basis of both the mass and collision cross section, providing information on their geometrical structures. If the mechanism proposed in ref 5 prevails, one would expect to observe a relatively mobile 248 Da bicyclic photofragment (1D in Figure 1) and perhaps a stabilized version of the 340 Da intermediate (1C in Figure 1). On the other hand, the

etinal protonated Schiff base (RPSB, Figure 1.1A) is a key molecule in the visual transduction cycle and in bacterial photosynthesis, where, in both cases, its action involves photoisomerization around bonds in the polyene chain. When embedded in a host protein, RPSB is remarkably photostable and can be repeatedly photoexcited without decomposing. However, in a vacuum environment, RPSB is more fragile, and exposure to visible light or collisions with neutral gas molecules causes its disintegration.1−5 Intriguingly, the dominant charged fragment at m/z = 248 Da cannot be generated through a simple chain cleavage and must result from a process involving cyclization of the polyene chain.5 The previously proposed decomposition mechanism,5 illustrated in Figure 1.1, involves a submillisecond rearrangement through a Diels−Alder mechanism, giving an energized, long-lived tricyclic structure that disintegrates on a millisecond time scale to produce a charged bicyclic 248 Da fragment and a toluene molecule (or an isomer thereof). Here, we consider an alternative mechanism for production of the 248 Da fragment from RPSB. As shown in Figure 1.2, a sequential electrocylization/fragmentation process leads directly to the β-ionone protonated Schiff base (Figure 1.2D) and a neutral toluene molecule. This process is analogous to the elimination of toluene and xylene from carotenes observed in mass spectrometric studies.6,7 Density functional theory (DFT) calculations at the M06-2X/cc-pVDZ level suggest that the fragment energies for the two mechanisms are comparable (−58 kJ/mol for 1D and −70 kJ/mol for 2E with respect to the energy of RPSB). To examine the RPSB fragmentation process, we use ion mobility mass spectrometry (IMMS) to probe the 248 Da © 2014 American Chemical Society

Received: July 7, 2014 Accepted: September 2, 2014 Published: September 2, 2014 3195

dx.doi.org/10.1021/jz501407n | J. Phys. Chem. Lett. 2014, 5, 3195−3199

The Journal of Physical Chemistry Letters

Letter

Figure 3. Laser-off ion count plotted against the mass (m/z) and cross section for collisions with N2 (σ). Red circles represent calculated data for isomers derived from the mechanism proposed in this work, whereas red squares correspond to isomers associated with the mechanism proposed by Toker et al.5 The ion count inside of the dashed rectangle has been multiplied by a factor of 10 for clarity. For expanded views of this plot, see the Supporting Information (SI).

Figure 1. RPSB fragmentation mechanisms proposed by Toker et al. (1)5 and in this work (2). In mechanism 1, RPSB undergoes a [4 + 2] cycloaddition to form a tricyclic intermediate structure, followed by loss of toluene. In mechanism 2, sequential 8π/6π electron electrocyclizations lead to an intermediate containing a fourmembered ring, followed by elimination of toluene through a 4π cycloreversion to give the protonated Schiff base of β-ionone. The two mechanisms involve loss of different sections of the polyene chain.

248 Da isomers with collision cross sections of 179.7, 176.8, and 167.5 Å2. Measured collision cross sections for the dominant 248 Da ions (179.7, 176.8 Å2) are closer to the predicted cross section for the monocyclic β-ionone PSB (2E; σcalc = 173.9 Å2 for the all-trans isomer) than the more compact, bicyclic structure proposed by Toker et al.5 (1D; σcalc = 166.3 Å2). Notably, we find no evidence for the stabilized tricyclic intermediate (1C), which is predicted to have a collision cross section 15% less than that of RPSB and, which, if it is formed rapidly and is long-lived, should be stabilized through collisions in our IMMS apparatus. Evidence for the formation of 2C and 2D is equivocal; predicted cross sections for the various stereoisomers of 2C and 2D (see the SI) are close to observed signal, although there is no exact correspondence. As demonstrated previously,21 when the drifting RPSB ions in our apparatus are exposed to light at low intensity (λ = 532 nm, I = 1−2 mJ/cm2/pulse), photodissociation is minimal, and instead, RPSB photoisomerizes from the all-trans form to various cis forms. However, at much higher intensities (I ≈ 30 mJ/cm2/pulse), we find that 248 Da photofragments are generated from RPSB. These photofragments have a very

sequential electrocyclization/fragmentation reaction should yield slower monocyclic β-ionone PSB photofragments (2E in Figure 1) and possibly the stabilized intermediates 2C and 2D. The ion count for electrosprayed RPSB solution is plotted as a function of mass and collision cross section with N2 in Figure 3. The ion population is dominated by 340 Da, σ = 227 Å2 ions associated with all-trans-RPSB but also includes 248 Da ions, as previously identified by Toker et al.5 These ions are produced in the source region of our apparatus (presumably through decomposition of RPSB ions in the electrospray desolvation capillary or first ion funnel). The ion mobility spectrum for the 248 Da ions (obtained with the mass filter tuned to 248 Da) is given in Figure 4a, clearly showing that there are in fact three

Figure 2. Ion mobility mass spectrometer for measuring collision cross sections of photofragments from RPSB. RPSB ions produced by an electrospray source pass through a heated desolvation capillary and into an ion funnel (IF1). The ions are injected from IF1 through an electrostatic gate into the drift region filled with N2 gas (P = 14 Torr) and subjected to an electric field of ∼44 V/cm. Here, they are exposed to the pulsed output of a frequency-doubled Nd:YAG laser (532 nm), causing photodissociation and production of 248 Da fragments. After passing through the drift tube, parent and fragment ions are collected by a second ion funnel (IF2), travel through an octopole ion guide, are mass selected by a quadrupole mass filter, and are sensed by an ion detector. 3196

dx.doi.org/10.1021/jz501407n | J. Phys. Chem. Lett. 2014, 5, 3195−3199

The Journal of Physical Chemistry Letters

Letter

Figure 5. Calculated PES for the thermal electrocyclization/ fragmentation cascade of the minimal model of RPSB. M06-2X/ccpVDZ electronic energies corrected for vibrational zero-point energy are in kJ/mol and are relative to the starting molecule 1. Further stereochemical details of the process are given in the SI.

The second electrocyclization has an activation energy of 77 kJ/ mol (TS2) and leads stereospecifically to a cis-bicyclo[4.2.0]octane framework (structure 3 in Figure 5). This cycloadduct is lower in energy by 34 kJ/mol compared to starting material 1. Subsequent dissociation through cycloreversion via TS3 occurs through a concerted bond cleavage (as confirmed by intrinsic reaction coordinate calculations) and produces toluene and a protonated Schiff base. Overall, the electrocyclization/fragmentation cascade is energetically very favorable, with the product association complex 4 being about 157 kJ/mol lower in energy than the starting material 1. Mechanism 2 is possible both thermally (as shown in Figure 5) and photochemically. The only possible difference lies in the stereoselectivity of the cyclization and fragmentation steps. However, as outlined below, we believe that photoexcited RPSB rapidly internally converts from the S1 to the S0 state so that the photoinduced electrocyclic reaction observed under our experimental conditions occurs for highly vibrationally excited ions on the ground-state potential energy surface (PES). Under these circumstances, the stereochemical progress of the 8πe and 6πe electrocylizations should follow courses appropriate for thermal conditions (conrotatory for the 8πe cyclization and disrotatory for the 6πe cyclization). Although, the sequential electrocyclizations should be highly stereospecific processes, apart from the cis fusion of the two rings, it is not possible to predict the relative stereochemistry at the cyclobutane ring for the full 2D (Figure 1) because the outcome of the electrocyclizations depends also on the geometry of the C9−C10 and CN double bond in RPSB. As explained above, under our experimental conditions, we cannot unequivocally determine whether 2C and 2D are sufficiently long-lived to enable their detection, and therefore, detailed knowledge of their stereochemistry is not essential at this stage. It should also be noted that DFT M06-2X/cc-pVDZ computations for a simplified model system for 1C did not reveal a fragmentation pathway leading to 1D. Instead, the calculations suggest that heterolytic cleavage of the C−NR+3 bond in 1C is a more viable process. The dominance of RPSB photoisomerization over photodissociation at low light levels causes us to believe that in our

Figure 4. Ion mobility spectra for 248 Da ions formed from (a) activation of RPSB in the ion source, (b) photodissociation of RPSB in the drift tube, and (c) electrosprayed protonated Schiff base of βionone. In each case, the peaks are fitted to three Gaussian functions constrained to have fwhm = 3 Å2. Calculated collision cross sections for isomers 1D and 2E (β-ionone) with N2 are shown in (a).

similar ion mobility spectrum to that of the 248 Da fragments formed in the ion source (Figure 4b). To confirm the identity of the 248 Da fragment, we synthesized β-ionone PSB (Figures 1 and 1.2E) and electrosprayed it directly into the IMMS apparatus. As shown in Figure 4c, the ion mobility spectrum for the electrosprayed sample corresponds unambiguously with the ion mobility spectra for the 248 Da fragments generated from RPSB in the ion source and through photodissociation from RPSB, conclusively identifying these fragments as β-ionone PSB. The three IMS peaks, which have slightly different relative intensities in Figure 4a−c, are presumably associated with βionone geometric isomers with different numbers of cis bonds along the polyene chain (calculated energies and predicted collision cross sections for β-ionone PSB isomers are given in the SI). At this stage, we assign the slowest peak to the all-trans isomer, the next fastest peak to molecules with a single cis linkage, and the fastest peak to isomers with a double cis linkage. The consecutive 8πe and 6πe electrocyclizations of retinal through mechanism 2 (Figure 1) are reminiscent of processes occurring in synthetic organic chemistry and in natural systems.22 To support the proposed rearrangement, we used DFT M06-2X/cc-pVDZ calculations to investigate a simplified RPSB model that mimics the C9−N polyene chain, replacing the β-ionone and N-butyl substituents by vinyl and methyl groups, respectively (Figure 5). The process involves several trans−cis isomerizations to achieve the required “curled” precursor conformation for the 8πe cyclization (1). The reaction then proceeds in a stepwise manner, whereby the first ring closure, which leads to an eight-membered ring (structure 2 in Figure 5) is associated with a barrier of 58 kJ/mol (TS1). 3197

dx.doi.org/10.1021/jz501407n | J. Phys. Chem. Lett. 2014, 5, 3195−3199

The Journal of Physical Chemistry Letters



COMPUTATIONAL METHODS To connect the measured collision cross sections with molecular structures, we determined equilibrium geometries for different 248 and 340 Da isomers using DFT calculations at the M06-2X/cc-pVDZ level.25 Collision cross sections were determined from the calculated equilibrium structures using the trajectory method as instituted in the MOBCAL program.26,27 Atom−atom potential energy parameters for the interaction between N2 and the colliding molecule were taken from ref 12. Calculated energies and collision cross sections for isomers of 1C, 1D, 2C, 2D, and 2E are given in the SI.

apparatus, photodissociation of RPSB involves consecutive absorption of two or more 532 nm photons.21 Possibly, the process occurs through absorption of one photon followed by rapid internal conversion, absorption of a second photon, again followed by internal conversion, trans−cis isomerizations, 8πe electrocylization, 6πe electrocylization, and dissociation, all on the ground-state PES. In contrast, RPSB dissociation in the ion storage ring is believed to involve absorption of a single photon.5 The magnitude of the calculated energy barriers for the simplified model suggests that single-photon photodissociation of RPSB should be possible over the 500−610 nm range, although the process may be sufficiently slow that in our experiment, the reactant RPSB ions are collisionally quenched in the drift tube, where the collision rate is ∼5 × 108 s−1 before surmounting the initial barrier to cyclization. On the other hand, in the ion ring experiment, where the pressure is ≤10−10 Torr and the collision rate is much lower,23 RPSB ions that have absorbed a single photon have much more time to rearrange and are able to dissociate on a millisecond time scale. In summary, we have used IMMS to probe the structure of charged molecular fragments from the chromophore RPSB in the gas phase and have unequivocally identified the 248 Da fragment as the protonated Schiff base of β-ionone. The proposed mechanism, which occurs after both thermal and photoexcitation, involves isomerization of the polyene chain, followed by a sequential electrocyclization/fragmentation cascade that leads to elimination of toluene. Eventually, the IMMS approach described in this Letter should be applicable to structural and mechanistic investigations of a broad range of gas-phase photochemical reactions, augmenting more cumbersome and expensive methods including isotopic labeling.



Letter



ASSOCIATED CONTENT

* Supporting Information S

Further experimental and computational information, including details for the synthesis of RPSB and β-ionone PSB, calculated energies and collision cross sections for intermediates and products shown in Figure 1, and Gaussian archive entries for structures in Figure 5. This material is available free of charge via the Internet at http://pubs.acs.org.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This research was supported under the Australian Research Council’s Discovery Project funding scheme (Project Numbers DP110100312 and DP120100100). The computations were supported by the National Computational Merit Allocation Scheme (Project m88). We thank Professor Matthew Bush for providing a modified version of the Mobcal code with N2 collision parameters described in ref 12, Luke Gamon for assistance with synthesis and collection of GC-MS data for the protonated Schiff base of β-ionone, and Sioe See Volaric for assistance with collection of LC-MS data.

EXPERIMENTAL METHODS

The ion mobility apparatus is the same as that used recently to investigate the photoisomerization of carbocyanine dyes and RPSB cations.19−21 In the current study, electrosprayed RPSB cations produced from a 10−5 M solution of trans-RPSB in 1:1 methanol/H2O (electrospray voltage of 3 kV; flow rate of 5 μL/min) were accumulated in an ion funnel before being launched in a 400 μs pulse into a 0.9 m drift tube containing N2 buffer gas (P = 14 Torr). The electric field in the drift tube (44 V/cm) was sustained by 90 ring electrodes. At the end of the drift tube, the ions were collected radially by a second ion funnel before passing through a 0.3 mm orifice into an octopole ion guide from which they exited through a 3 mm orifice into a quadrupole mass filter. Ions were sensed by a channeltron detector connected to a discriminator and a multichannel scaler. The apparatus was run at 20 Hz, with alternate ion packets exposed to the 532 nm output from a pulsed frequencydoubled Q-switched Nd:YAG laser. The ions’ arrival time distribution (ATD) was built up as a histogram of ion counts versus time. The mobility resolution of the instrument was typically td/Δtd = 60, which can be compared to a maximum, diffusion-limited resolution under the prevailing conditions of 120.19 Under typical operating conditions with a drift field of 44 V/cm and N2 buffer gas pressure of 14 Torr, the effective temperature of the drifting RPSB ions is predicted to be ∼300 K.24



REFERENCES

(1) Andersen, L. H.; Nielsen, I. B.; Kristensen, M. B.; El Ghazaly, M. O. A.; Haacke, S.; Nielsen, M. B.; Petersen, M. Å. Absorption of SchiffBase Retinal Chromophores in Vacuo. J. Am. Chem. Soc. 2005, 127, 12347−12350. (2) Nielsen, I.; Lammich, L.; Andersen, L. S1 and S2 Excited States of Gas-Phase Schiff-Base Retinal Chromophores. Phys. Rev. Lett. 2006, 96, 018304. (3) Lammich, L.; Nielsen, I. B.; Sand, H.; Svendsen, A.; Andersen, L. H. Probing the Sub-Microsecond Photodissociation Dynamics in GasPhase Retinal Chromophores. J. Phys. Chem. A 2007, 111, 4567−4572. (4) Rajput, J.; Rahbek, D. B.; Andersen, L. H.; Hirshfeld, A.; Sheves, M.; Altoè, P.; Orlandi, G.; Garavelli, M. Probing and Modeling the Absorption of Retinal Protein Chromophores in Vacuo. Angew. Chem., Int. Ed. 2010, 49, 1790−1793. (5) Toker, Y.; Rahbek, D. B.; Kiefer, H. V.; Rajput, J.; Antoine, R.; Dugourd, P.; Nielsen, S. B.; Bochenkova, A. V.; Andersen, L. H. Photoresponse of the Protonated Schiff-Base Retinal Chromophore in the Gas Phase. Phys. Chem. Chem. Phys. 2013, 15, 19566−19569. (6) Carnevale, J.; Cole, E. R.; Nelson, D.; Shannon, J. S. Chemical Ionisation Mass Spectrometry of Carotenoids. Biomed. Mass Spectrom. 1978, 5, 641−646. (7) van Breemen, R. B.; Dong, L.; Pajkovic, N. D. Atmospheric Pressure Chemical Ionization Tandem Mass Spectrometry of Carotenoids. Int. J. Mass Spectrom. 2012, 312, 163−172. 3198

dx.doi.org/10.1021/jz501407n | J. Phys. Chem. Lett. 2014, 5, 3195−3199

The Journal of Physical Chemistry Letters

Letter

(8) Lanucara, F.; Holman, S. W.; Gray, C. J.; Eyers, C. E. The Power of Ion Mobility-Mass Spectrometry for Structural Characterization and the Study of Conformational Dynamics. Nat. Chem. 2014, 6, 281−294. (9) Kanu, A. B.; Dwivedi, P.; Tam, M.; Matz, L.; Hill, H. H., Jr. Ion Mobility-Mass Spectrometry. J. Mass Spectrom. 2008, 43, 1. (10) Wyttenbach, T.; Bowers, M. T. Intermolecular Interactions in Biomolecular Systems Examined by Mass Spectrometry. Annu. Rev. Phys. Chem. 2007, 58, 511−533. (11) Pierson, N. A.; Chen, L.; Russell, D. H.; Clemmer, D. E. Cis− Trans Isomerizations of Proline Residues Are Key to Bradykinin Conformations. J. Am. Chem. Soc. 2013, 135, 3186−3192. (12) Campuzano, I.; Bush, M. F.; Robinson, C. V.; Beaumont, C.; Richardson, K.; Kim, H.; Kim, H. I. Structural Characterization of Drug-Like Compounds by Ion Mobility Mass Spectrometry: Comparison of Theoretical and Experimentally Derived Nitrogen Collision Cross Sections. Anal. Chem. 2012, 84, 1026−1033. (13) Shvartsburg, A. A.; Smith, R. D. Fundamentals of Traveling Wave Ion Mobility Spectrometry. Anal. Chem. 2008, 80, 9689−9699. (14) von Helden, G.; Gotts, N. G.; Bowers, M. T. Experimental Evidence for the Formation of Fullerenes by Collisional Heating of Carbon Rings in the Gas Phase. Nature 1993, 363, 60−63. (15) Hudgins, R. R.; Imai, M.; Jarrold, M. F.; Dugourd, P. HighResolution Ion Mobility Measurements for Silicon Cluster Anions and Cations. J. Chem. Phys. 1999, 111, 7865−7870. (16) Both, P.; Green, A. P.; Gray, C. J.; Šardzk, R.; Voglmeir, J.; Fontana, C.; Austeri, M.; Rejzek, M.; Richardson, D.; Field, R. A.; et al. Discrimination of Epimeric Glycans and Glycopeptides using IM-MS and Its Potential for Carbohydrate Sequencing. Nat. Chem. 2013, 6, 65−74. (17) Wu, C.; Siems, W. F.; Klasmeier, J.; Hill, H. H., Jr. Separation of Isomeric Peptides Using Electrospray Ionization/High-Resolution Ion Mobility Spectrometry. Anal. Chem. 2000, 72, 391−395. (18) Pringle, S. D.; Giles, K.; Wildgoose, J. L.; Williams, J. P.; Slade, S. E.; Thalassinos, K.; Bateman, R. H.; Bowers, M. T.; Scrivens, J. H. An Investigation of the Mobility Separation of Some Peptide and Protein Ions Using a New Hybrid Quadrupole/Travelling Wave IMS/ OA-TOF Instrument. Int. J. Mass Spectrom. 2007, 261, 1−12. (19) Adamson, B. D.; Coughlan, N. J. A.; Continetti, R.; Bieske, E. J. Changing the Shape of Molecular Ions: Photoisomerization Action Spectroscopy in the Gas Phase. Phys. Chem. Chem. Phys. 2013, 15, 9540−9548. (20) Adamson, B. D.; Coughlan, N. J. A.; da Silva, G.; Bieske, E. J. Photoisomerization Action Spectroscopy of the Carbocyanine Dye DTC+ in the Gas Phase. J. Phys. Chem. A 2013, 117, 13319−13325. (21) Coughlan, N. J. A.; Catani, K. J.; Adamson, B. D.; Wille, U.; Bieske, E. J. Photoisomerization Action Spectrum of Retinal Protonated Schiff Base in the Gas Phase. J. Chem. Phys. 2014, 140, 164307. (22) Nicolaou, K. C.; Petasis, N. A.; Zipkin, R. E. The Endiandric Acid Cascade. Electrocyclizations in Organic Synthesis. 4. “Biomimetic” Approach to Endiandric Acids A−G. Total Synthesis and Thermal Studies. J. Am. Chem. Soc. 1982, 104, 5560−5562. (23) Andersen, J. U.; Hvelplund, P.; Nielsen, S. B.; Tomita, S.; Wahlgreen, H.; Møller, S. P.; Pedersen, U. V.; Forster, J. S.; Jørgensen, T. J. D. The Combination of an Electrospray Ion Source and an Electrostatic Storage Ring for Lifetime and Spectroscopy Experiments on Biomolecules. Rev. Sci. Instrum. 2002, 73, 1284−1287. (24) Revercomb, H. E.; Mason, E. A. Theory of Plasma Chromatography/Gaseous Electrophoresis. Anal. Chem. 1975, 47, 970−983. (25) Frisch, M. J.; et al. Gaussian 09, revision A.1; Gaussian, Inc.: Wallingsford, CT, 2009. (26) Mesleh, M.; Hunter, J.; Shvartsburg, A.; Schatz, G.; Jarrold, M. Structural Information from Ion Mobility Measurements: Effects of the Long-Range Potential. J. Phys. Chem. 1996, 100, 16082−16086. (27) Shvartsburg, A. A.; Jarrold, M. F. An Exact Hard-Spheres Scattering Model for the Mobilities of Polyatomic Ions. Chem. Phys. Lett. 1996, 261, 86−91.

3199

dx.doi.org/10.1021/jz501407n | J. Phys. Chem. Lett. 2014, 5, 3195−3199